Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Journal of

Clinical Medicine

Review
Myasthenia Gravis: Epidemiology, Pathophysiology and
Clinical Manifestations
Laura Dresser * , Richard Wlodarski, Kourosh Rezania and Betty Soliven

Department of Neurology, University of Chicago, Chicago, IL 60637, USA;


richard.wlodarski@uchospitals.edu (R.W.); krezania@neurology.bsd.uchicago.edu (K.R.);
bsoliven@neurology.bsd.uchicago.edu (B.S.)
* Correspondence: laura.dresser@uchospitals.edu

Abstract: Myasthenia gravis (MG) is an autoimmune neurological disorder characterized by defective


transmission at the neuromuscular junction. The incidence of the disease is 4.1 to 30 cases per
million person-years, and the prevalence rate ranges from 150 to 200 cases per million. MG is
considered a classic example of antibody-mediated autoimmune disease. Most patients with MG have
autoantibodies against the acetylcholine receptors (AChRs). Less commonly identified autoantibodies
include those targeted to muscle-specific kinase (MuSK), low-density lipoprotein receptor-related
protein 4 (Lrp4), and agrin. These autoantibodies disrupt cholinergic transmission between nerve
terminals and muscle fibers by causing downregulation, destruction, functional blocking of AChRs,
or disrupting the clustering of AChRs in the postsynaptic membrane. The core clinical manifestation
of MG is fatigable muscle weakness, which may affect ocular, bulbar, respiratory and limb muscles.
Clinical manifestations vary according to the type of autoantibody, and whether a thymoma is present.

 Keywords: myasthenia gravis; acetylcholine receptor; autoantibodies; cytokines; B cells; T cells
Citation: Dresser, L.; Wlodarski, R.;
Rezania, K.; Soliven, B. Myasthenia
Gravis: Epidemiology,
Pathophysiology and Clinical 1. Introduction
Manifestations. J. Clin. Med. 2021, 10,
Myasthenia gravis (MG) is the most common autoimmune disorder that affects the
2235. https://doi.org/10.3390/
neuromuscular junction. MG is largely a treatable disease but can result in significant
jcm10112235
morbidity and even mortality. This can usually be prevented with a timely diagnosis and
appropriate treatment of the disease. MG is a heterogeneous disease from a phenotypic
Academic Editor: Cristoforo Comi
and pathogenesis standpoint. The spectrum of symptoms ranges from a purely ocular
form to severe weakness of the limb, bulbar and respiratory muscles. The age of onset is
Received: 27 April 2021
Accepted: 17 May 2021
variable from childhood to late adulthood with disease peaks in younger adult women and
Published: 21 May 2021
older men [1].
MG is considered a classic example of antibody-mediated autoimmune disease. It can
Publisher’s Note: MDPI stays neutral
also be viewed as an example of a class II hypersensitivity reaction, as IgG autoantibodies
with regard to jurisdictional claims in
react with intra or extracellular antigens, leading to end-organ damage. Most patients with
published maps and institutional affil- MG have autoantibodies against the acetylcholine receptors (AChRs) [2,3], and a minority
iations. are seropositive for antibodies directed to muscle-specific kinase (MuSK) [4,5], low-density
lipoprotein receptor-related protein 4 (Lrp4) [6,7] or agrin [8,9]. These antibodies also
provide the basis for defining disease subgroups and help delineate phenotypic variants. In
a subgroup of MG patients, striational antibodies have also been identified, which include
Copyright: © 2021 by the authors.
antibodies against titin, ryanodine receptor, and the alpha subunit of the voltage-gated K+
Licensee MDPI, Basel, Switzerland.
channel (Kv1.4). These antibodies mostly serve as biomarkers of disease severity and are
This article is an open access article
often detected in patients with late-onset MG or with thymoma, and some of them have
distributed under the terms and concomitant myositis and/or myocarditis [10,11].
conditions of the Creative Commons Although MG is mediated by autoantibodies, different subtypes of T cells and their
Attribution (CC BY) license (https:// cytokines also play important roles in the pathogenesis. In this review, we briefly discuss
creativecommons.org/licenses/by/ the epidemiology, clinical manifestations, and genetic predisposing factors of adult MG,
4.0/). then provide an overview of pediatric MG, followed by an update on MG pathophysiology.

J. Clin. Med. 2021, 10, 2235. https://doi.org/10.3390/jcm10112235 https://www.mdpi.com/journal/jcm


J. Clin. Med. 2021, 10, 2235 2 of 17

2. Epidemiology
MG is a rare neurological disease and pediatric MG is even more uncommon. Both
incidence and prevalence have significant geographical variations, but it is believed that
MG incidence has increased worldwide over the past seven decades. The prevalence
of MG was estimated at 1 in 200,000 from 1915 to 1934, increased to 1 per 20,000 after
the introduction of anticholinesterase drugs in 1934, and rose to 1 per 17,000 population
after the discovery of AChR antibodies in 1969 [12]. Prevalence rates range from 150 to
200 cases per million, and they have steadily increased over the past 50 years, at least partly
due to improvements in recognition, diagnosis, treatment, and an overall increase in life
expectancy [13]. More recent studies addressing incidence rates have been conducted in
Europe and show a wide range from 4.1 to 30 cases per million person-years [14,15].
The annual rate is lower in studies coming from North America and Japan, with the
incidence ranging from 3 to 9.1 cases per million [1]. Lower incidence and prevalence
rates have been reported in a large study from China at 0.155–0.366 per million, and
2.19–11.07 per 100,000, respectively [16]. Two population-based studies from Korea showed
a prevalence of 9.67–10.42 per 100,000 people in 2010, which increased to 12.99 per 100,000
in 2014 [17,18]. On the other hand, a smaller study using records of a hospital-based Health
Maintenance Organization estimated an incidence of MG at 38.8 per 1,000,000 person-years
for the Argentinian population [19]. Different study methodologies, including diagnostic
criteria and other sources of bias, such as the small size of the study population and the
underestimation of patients with milder disease, likely play a factor in the significant
variability of incidence rates over time and across different geographical regions.
Incidence rates have a bimodal distribution in women, with peaks around age 30 and
50. In men, the incidence increases steadily with age and with the highest rates between
age 60 and 89 [20]. Women are more commonly affected before age 40, with a female: male
ratio of 3:1 for early-onset MG. In the fifth decade of life, women and men are equally
affected, while men have a higher proportion after age 50, with a male: female ratio of
3:2 [21]. Around 10% of cases are pediatric, which is defined as onset before age 18 [13].
MG can affect people of all race and ethnic backgrounds and is slightly more prevalent in
patients of African ancestry [22–24]. Furthermore, MG phenotype may vary depending
on the ethnic background. In a retrospective study from South Africa, black patients were
more likely to have treatment-resistant ophthalmoplegia and ptosis than whites, whereas
the whites were more likely to develop treatment refractory generalized MG [25]. The age
at diagnosis was 17 years higher in Caucasians than non-Caucasians in another cohort
of patients with ocular MG [26]. In a US study, Oh et al. found that MG started earlier
and had a more severe phenotype in African Americans than in Caucasians [22]. The
seronegative African Americans had a higher percent of MuSK seropositivity in that study
(50% vs. 17% in the whites). On the other hand, patients of Asian ancestry have higher rates
of MuSK antibodies compared to Caucasians and individuals of African ancestry [27,28].
MuSK-associated MG is also more prevalent among those living in latitudes closer to the
equator [22,29].
The mortality rate of MG has dramatically declined from the early 20th century after
the availability of acetylcholine esterase inhibitors, immunosuppressants, intravenous
immunoglobulin and advanced respiratory care. However, the mortality rate from the
disease remains at 5–9%, being slightly higher in males than females [12]. Using the
US Nationwide Inpatient Sample (NIS) database for the years 2000 to 2005, the overall
in-hospital mortality rate was estimated as 2.2%, but higher in those with MG crisis
(4.7%), with the main predictors of death being older age and the presence of respiratory
failure [24].
J. Clin. Med. 2021, 10, 2235 3 of 17

3. Subtypes of MG and Their Clinical Manifestations


3.1. MG Due to Antibodies against AChR (AChR-MG)
3.1.1. Effector Mechanisms
Nicotinic AChR is a heteropentamer consisting of two α-subunits and one each of β-,
δ-subunit, and γ-subunit (embryonic type) or ε-subunit (adult type), which are organized
around a central pore [30]. Antibodies against the AChR are found in approximately
80% of MG patients. At least half of the AChR autoantibodies are directed at the AChR
α-subunits [31]. They are believed to be more pathogenic than those directed against the
beta subunit [32]. This is likely due to the location of the alpha subunit within the receptor,
which leaves it more exposed to antibodies, as well as its role in modulating the receptor
sensitivity to ACh binding [33,34]. In addition, there are two alpha subunits per receptor.
AChR antibodies are predominantly of the IgG1 and IgG3 subclasses [35]. IgG2 and
IgG4 subclasses are also identified, but in fewer cases [36]. The pathogenic mechanisms
and functional spectrum of AChR antibodies are varied, but overall, they impair receptor
function by either binding, blocking, or modulating its activity. The predominant mech-
anism is the binding of the antibody and activation of the complement cascade, leading
to the formation of the membrane attack complex (MAC), which causes damage of the
postsynaptic membrane and destruction of synaptic folds which contain AChRs and as-
sociated proteins, including voltage-gated sodium channels [35]. Other mechanisms of
pathogenicity include: (1) antigenic modulation by the binding and crosslinking of AChRs,
leading to increased endocytosis and degradation [37]; and (2) the impairment of AChR
function, either by the blocking of ACh binding to the receptor [38] or the prevention
of channel opening [39]. Most blocking antibodies appear in association with binding
antibodies, and only rarely are they unique. Therefore, this final mechanism is believed
to be rare and its clinical implications are not clear [40]. However, the administration of
blocking antibodies causes an acute and severe weakness in rodents [41].

3.1.2. Clinical Manifestations


The core clinical manifestation of MG is fatigable muscle weakness, worsened by exer-
tion and improved by rest. The most common presenting symptoms are ocular, with double
vision and ptosis. Most patients will develop diplopia and/or ptosis some time during the
course of their disease. In addition, up to 80% of patients with ocular onset will go on to
develop generalized symptoms, usually within two years of disease onset [42,43]. A recent
population-based study conducted by the Mayo Clinic found that 51% of patients presented
with ocular onset, and 55% of these developed generalized symptoms [44]. Bulbar muscles
are also frequently involved, resulting in flaccid dysarthria, dysphagia, and facial and jaw
weakness [45]. Axial weakness can also be present, with neck flexion weakness usually
more common than weakness of neck extension. In a previous retrospective study from
our center, about 10% of MG patients developed head-drop some time during the disease
course [46]. Head-drop was associated with age >60 and male patients in that study. Limb
muscle weakness tends to be symmetric and proximal, and patients commonly complain
of difficulty climbing stairs, getting up from chairs, and raising arms above their head. In
some instances, distal muscles can be predominantly affected, either in a symmetric or
asymmetric distribution. For example, some patients complain of weakness in finger and
wrist extension and flexion, as well as foot drop. Finally, 15–20% of patients with AChR-
MG can develop respiratory weakness requiring mechanical ventilation (MG crisis) [21].
Spontaneous remissions for different lengths of time may occur in the course of adult-onset
MG. In an earlier study conducted before the widespread use of steroids and other immuno-
suppressants, approximately one fourth of the patients had a complete or near complete
spontaneous remission, lasting an average of 4.6 years and up to 17 years [47]. Half of the
remissions occurred during the first year after onset. A study of Oosterhuis et al. found
that 22% of patients treated with anticholinesterase medications only had spontaneous
remission [42]. The remission lasted more than 12 months in duration in half of those
patients, with the maximum duration of the remission being 6 years in that study.
J. Clin. Med. 2021, 10, 2235 4 of 17

3.1.3. AChR MG Subtypes


Ocular MG
Most MG patients with ocular symptoms at onset will progress to generalized forms
of the disease, usually within two years of onset. Of the remaining, 90% will continue
to have ocular manifestations only. Hence, ocular MG is defined by isolated extra-ocular
involvement for a period of ≥2 years. Over half of the patients in this group have antibodies
against AChRs [21]. There are several explanations as to why extraocular muscles (EOMs)
are preferentially affected in MG. EOMs fatigue easily, as they require tonic contractions to
sustain gaze in a specific direction, and fibers have a high frequency of synaptic firing, and
develop tension faster [48]. In addition, EOMs have a lower density of AChR, thus making
them more susceptible to symptoms. It is also theorized that differing epitope expression
in EOMs plays a role in their preferential involvement [49].

Generalized AChR Ab Positive MG (AChR-MG): Early vs. Late Onset


Early-onset MG (EOMG) corresponds to patients presenting before age 50. There is a
female preponderance, with a female to male ratio of 3:1. Patients in this category have
a higher incidence of thymic hyperplasia, and thymectomy has been proven effective in
improving clinical outcomes and minimizing the need for immunotherapy [50]. Late-onset
MG (LOMG) is defined by onset after the age of 50. There is no female predominance in
this group; on the contrary, there can be a slightly higher prevalence among men, especially
after age 60. Thymic hyperplasia is rare and response to thymectomy is poorer [45].
There is a higher prevalence of autoimmune disease among family members and a
high correlation with HLA-DR and HLA-B8 haplotypes [51]. There is frequent association
of A1-B8-DR3 haplotype with early-onset MG [52]. In a study on non-thymomatous AChR
Ab + LOMG of Italian ancestry, Spagni et al. found a positive association with HLA-
DRB1*07 and HLA-DQB1*02, whereas HLA-DRB1*02, HLA-DRB1*03, HLA-DRB1*11, and
HLA-DQB1*03 were the protective alleles [53]. A study comparing cohorts of patients with
EOMG, LOMG and MuSK-MG in a single center in Turkey found a strong association for
class I HLA-B/MICA in EOMG patients, specifically HLA-B*08:01. On the other hand, no
association was found between LOMG and HLA class I, but it detected an association with
HLA-DQA1 and HLA-DRB1 [54]. Another large study on Norwegian MG patients older
than 60 showed a strong association with HLA-DRB1*15:01 [55].
Genetic variations within other loci have been associated with predisposition to
AChR-MG. A large genome-wide association study involving patients from North America
and Italy identified different haplotypes across the HLA region, cytotoxic T-lymphocyte-
associated protein 4 gene (CTLA4) and tumor necrosis factor receptor 4 superfamily,
member 11a, (TNFRSF11A) and NFκB activator genes in early- and late-onset MG cases [56].
Variations in the CTLA4 and HLA-DQA1 loci were associated with both early- and late-
onset cases, whereas genetic variation within the TNFSRF11A locus was a susceptibility
factor only in the late-onset cases in that study.

Thymoma-Associated MG
MG is the most common paraneoplastic disorder associated with thymoma. Other
thymoma-related disorders with lower association rates include myositis, Morvan syn-
drome and pure red aplasia. About 50% of patients with a thymoma develop positive
AChR antibodies without clinical manifestations, and approximately 30% will develop
MG [57]. Conversely, 10–20% of patients with MG have thymomas [58]. Response to
thymectomy is variable, usually worse than in patients with EOMG [50]. Studies on HLA
alleles did not reveal a consistent association between HLA and thymomatous MG [59].

3.2. MuSK Antibody-Associated MG (MuSK-MG)


3.2.1. Effector Mechanisms
MuSK is a membrane protein that is critical to the clustering of AChRs in the neuro-
muscular junction. Agrin, which is secreted from the presynaptic terminal, interacts with
J. Clin. Med. 2021, 10, 2235 5 of 17

Lrp4, resulting in the reorientation of the Lrp4/MuSK complex, which in turn leads to
the activation of MuSK through its phosphorylation. Phosphorylated MuSK activates a
downstream signaling pathway that leads to the clustering of AChRs (Figure 1) [60].

Figure 1. Diagram depicting the secretion of agrin from the presynaptic membrane and its interaction
with Lrp4, which results in reorganization and reorientation of MuSK, promoting a signaling pathway
that leads to synaptic differentiation, including clustering of AChRs. This involves recruiting rapsyn
which links AChRs to the cytoskeleton (not shown).

Antibodies against MuSK are found in approximately 7–10% of all MG patients and
up to 40% of patients with generalized MG who are seronegative for AChR Abs. There is a
female predominance, with up to 85% of MuSK positive patients being female [61]. Animal
studies demonstrated that MuSK antibodies are pathogenic, as they cause severe weakness
when administered to healthy mice [62]. In contrast to AChR antibodies, MuSK antibody
titers correlate with disease severity [63]. The concurrence of seropositivity for both AChR
and MuSK antibodies has rarely been reported [64], but in general these are considered
distinct entities.
MuSK antibodies belong mainly to the IgG4 subclass. They do not fix complement and
are not strong activators of cell-mediated cytotoxicity [65]. Given the unique ability of IgG4
to undergo Fab arm exchange, MuSK antibodies are functionally monovalent, and they
cannot crosslink antigens of the same class. The mechanism by which MuSK antibodies
exert their pathogenic effect on the neuromuscular junction is via binding to the Ig-like
domain of the protein, preventing its phosphorylation, and subsequently disrupting the
Agrin-Lrp4-MuSK-Dok-7 signaling pathway. Dok-7 is a muscle-intrinsic activator of MuSK
and it is required for synaptogenesis [66]. This ultimately causes a reduction in the density
of AChRs in the postsynaptic membrane [67,68].

3.2.2. Clinical Manifestations


MuSK-MG predominantly affects young adults, and is more prevalent among patients
of African descent and those living close to the equator in European and Asian nations [45].
This is likely due to a genetic predisposition and not to environmental factors. Muscle
weakness preferentially affects cranial and bulbar muscles. Over 40% of patients present
with bulbar weakness, usually associated with neck and respiratory involvement [69,70].
Some patients have tongue atrophy. About 30% of patients present with diplopia and/or
J. Clin. Med. 2021, 10, 2235 6 of 17

ptosis. Limb weakness can be uncommon, but when present it tends to be severe and
associated with muscle atrophy. Diurnal variations in strength are less common. There
is no clear association between MuSK-associated MG and thymic pathology [21]. From
the genetic predisposition standpoint, MuSK-associated MG has been associated with
DQB1*05 and HLA-DRB1*14/DRB1*16 [71,72].
Patients who are seropositive for both AChR and MuSK are rare, and it is not certain
if they should be categorized as an MG subtype. In a study from southern China, Zhang
et al. demonstrated that the phenotype of double-seropositive patients is more severe than
AChR-MG and more similar to the MuSK-associated MG [73].

3.3. Double-Seronegative Generalized MG


This is a heterogenous group of patients who share negative results for AChR and
MuSK antibody testing. Cell-based assays can detect lower titers of these antibodies in
patients previously reported as double-negative by more common serologic assays. It is
likely that this subgroup of patients has antibodies against antigens not yet identified [51].
In general, these patients present similar to those positive for AChR antibodies in regard to
the distribution of muscle weakness, severity, and response to treatment. Cortes-Vicente
and colleagues reported a cohort of double-seronegative MG patients who presented
mainly with milder forms of disease at onset, less bulbar involvement, and younger age at
presentation when compared to AChR-MG [74].

3.4. Lrp4 Antibody-Associated MG (Lrp4-MG)


3.4.1. Effector Mechanisms
Lrp4 is the postsynaptic receptor of nerve-derived agrin. The binding of agrin to Lrp4
activates MuSK and initiates a cascade of events leading to the aggregation of AChRs in the
neuromuscular junction. Lrp4 antibodies are present in 2–50% of double seronegative MG
cases [6,75,76]. Many Lrp4-MG patients also have antibodies against agrin. Lrp4 antibodies
are mostly of the IgG1/IgG2 subclass and are believed to be directly pathogenic by disrupt-
ing the activation of MuSK [75]. In patients with congenital myasthenic syndromes due
to gene mutations in agrin, neuromuscular junctions are less stable and AChRs are more
dispersed and un-clustered. Given these findings, the most likely pathologic mechanism of
anti-Lrp4 antibodies is a reduction in AChR clustering. Other mechanisms are possible as
well, as IgG1 can activate complement [6,76].

3.4.2. Clinical Manifestations


There is a female predominance, and age at onset is variable, but patients tend to
present before age 50. Patients can present with ocular or generalized MG, but it is believed
that symptom severity is overall milder in this subgroup. About 20% have only ocular
symptoms after 2 years of disease [75]. However, patients with a combination of anti-
Lrp4 and anti-agrin antibodies may have more severe symptoms [75]. Of note, Lrp4
Ab has also been reported in other neurological diseases, including amyotrophic lateral
sclerosis [77,78]; thus, further studies are required to validate its specificity in the diagnosis
of MG. The predisposing genetic markers for Lrp4 and double-seronegative MG have not
been reported.

4. Pediatric MG
Juvenile MG (JMG) is defined as MG in patients younger than 18 years of age. A
meta-analysis of epidemiological studies estimated the incidence of JMG between 1 and
5 cases per million person-years [1]. JMG is, however, more prevalent in the Asian than
the European populations. For example, Taiwanese and Japanese studies showed inci-
dence rates ranging from 3.7 to 8.9 per million person-years [1,79]. In other studies on
the Asian population, JMG constituted up to 50% of total MG cases, mostly of ocular
type [80,81]. There is also a higher prevalence in patients of African descent compared to
J. Clin. Med. 2021, 10, 2235 7 of 17

Caucasians [82,83]. Pediatric cases constituted 10–15% of total MG cases in a US study [13].
Similar to the young adults, JMG has a female preponderance [84,85].
A total of 16–38% of JMG cases are ocular [83,84]. Ocular JMG is more common in
pre-pubertal children, and post-pubertal children tend to have a greater proportion of
generalized MG [84–87]. As ocular symptoms often occur during critical times during
a child’s development, there is a risk for long-term sequelae such as strabismus and
amblyopia if this condition is not treated aggressively [88]. The generalization of symptoms
occurs in 20–25% of JMG patients [83,88], much lower than the rate of generalization in
adults, which can be up to 80% [12]. The course of generalized JMG is unpredictable and
may be associated with recurrent exacerbations including myasthenic crisis. On the other
hand, JMG patients may undergo periodic spontaneous remissions, at a higher rate than
the adults, often lasting for years [83,89].
Similar to adult-onset MG, the most common pathogenic antibodies detected in JMG
are AChR Ab, followed by MuSK and Lrp4. Fifty-three to 86% of generalized and ~40% of
ocular JMG patients are seropositive for AChR Abs [83,84,90]. It should be noted that ~40%
of patients seronegative for AChR Ab may become seropositive within 2 years of follow
up [89]. Up to 40% of JMG patients who are seronegative for AChR Ab are seropositive for
MuSK Abs [83,91]. Similar to adult-onset MG, anti-MuSK Ab seropositivity is predictive of
a more severe phenotype often associated with respiratory and bulbar muscle weakness
and atrophy. JMG may also be associated with anti-Lrp4 Ab, which manifests with a milder,
predominantly ocular phenotype [91,92]. In a large retrospective study from China, 13 of
455 (2.9%) patients who were seronegative for AChR and MuSK Abs had Lrp4 antibodies;
53.8% of them were children [92]. The pathogenesis of JMG is similar to the adult cases
and is described above.
JMG is a separate entity from transient neonatal MG, which occurs in 10–15% of
neonates born to mothers with MG as a result of the passive transfer of maternal antibodies
in utero [93,94]. While neonatal MG usually resolves spontaneously over weeks to months,
the affected infants may have generalized hypotonia, respiratory distress, poor suck in
addition to extraocular muscle weakness, necessitating the use of respiratory support,
treatment with neostigmine and, in severe cases, plasma exchange [93,94]. Furthermore,
the transfer of antibodies directed to fetal AChRs may result in persistent myopathic
features (i.e., fetal AChR inactivation syndrome), with symptoms ranging from mild facial
and bulbar weakness in the affected infants to arthrogryposis multiplex congenita [95].
Other JMG mimics include congenital myasthenic syndrome, mitochondrial diseases,
demyelinating polyneuropathies, and congenital myopathies. The asymmetric nature
of ptosis and ophthalmoparesis in ocular JMG may help distinguish it from congenital
myasthenic syndromes and mitochondrial diseases, as the extraocular weakness and
ptosis in the latter are usually, but not always, symmetrical [96]. Congenital myasthenic
syndromes are a particularly heterogeneous group of disorders and are beyond the scope
of this chapter.

5. Pathophysiology
5.1. Physiology and Organization of the Neuromuscular Junction
The neuromuscular junction is the site of impulse transmission between nerve ter-
minals and muscle fibers. This process requires the release of presynaptic acetylcholine
(ACh) and its subsequent binding to a postsynaptic ACh receptor. Synaptic vesicles con-
taining ACh are released from the presynaptic membrane after an action potential activates
voltage-gated calcium channels, allowing an influx of calcium into the nerve terminal [97].
The diffusion time of ACh through the synaptic cleft is very short, and it is modulated
by the enzyme ACh esterase (AChE), which promotes ACh degradation. The sponta-
neous release of synaptic vesicles generates the so-called miniature end plate potential
(MEPP), while upon nerve fiber stimulation/depolarization, a synchronous release of many
synaptic vesicles causes large depolarization of the end plate membrane, generating an
evoked endplate potential (EPP) (Figure 2). This, in turn, triggers an action potential in the
J. Clin. Med. 2021, 10, 2235 8 of 17

myofiber, which ultimately leads to its contraction. The amount of ACh released into the
synapse is usually higher than that required to generate an action potential, which allows
for reliable transmission [97,98]. The binding of ACh to its receptors in the postsynaptic
membrane opens the ACh cation-specific channel, leading to localized depolarization and
activation of adjacent voltage-gated sodium channels. This allows for the translation of the
chemical reaction into an electric signal, which is the muscle fiber action potential. The role
of AChE in the hydrolyzation of ACh is therefore crucial, as it prevents a single molecule
of ACh from repetitively activating AChRs. The effectiveness of neuromuscular junction
transmission is also determined by the amount of ACh released by the nerve terminal, the
density of ACh receptors in the postsynaptic membrane, and the density of voltage-gated
sodium channels at the endplate. The latter is dependent on the presence of folds in the
postsynaptic membrane. These folds determine the density of the voltage-gated sodium
channels in the postsynaptic membrane, and therefore increase the efficient coupling of the
localized EPP to the myofiber action potential [98].
Neuromuscular junction disorders such as MG disrupt the cascade of events that
lead to reliable muscle contraction. In addition, there is a reduction in the number of
AChRs and voltage-gated sodium channels as the result of complement-related injury to
the postsynaptic membrane in MG. The resultant inefficient neuromuscular transmission is
reflected in decremental response in the amplitude of compound muscle action potential
(CMAP) during repetitive nerve stimulation (RNS) and abnormal jitter or blocking in single
fiber EMG [98,99].

Figure 2. Neuromuscular transmission in normal individuals (A) and in patients with MG (B). Decreased density of the
AChR and complement-mediated damage to the postsynaptic membrane in MG patients result in decrease in miniature end
plate potential (MEPP), which occurs with spontaneous release of AChR vesicles, as well as endplate potential (EPP) in
response to nerve action potential of the presynaptic membrane. Diminished amplitude of EPP in MG results in impaired
neuromuscular transmission.

5.2. Immune Dysregulation in MG


Defective B Cell Tolerance
B cell tolerance is mediated by clonal deletion or receptor editing in newly generated
B cell clones in the bone marrow when they reach the stage of immature B cells. A second
checkpoint occurs on the new emigrant/transitional B cells before they enter the mature
naïve B cell compartment. Lee et al. found that the frequency of new emigrant/transitional
J. Clin. Med. 2021, 10, 2235 9 of 17

B cells and mature B cells that express polyreactive and autoreactive B cell receptors (BCRs)
is higher in both AChR-MG and MuSK-MG, which would support the concept that patients
with MG have defects in both central and peripheral B cell tolerance (Figure 3) [100]. As
a result, these patients are also at higher risk of developing other autoimmune diseases
such as systemic lupus erythematosus, rheumatoid arthritis, and thyroiditis. A break in
tolerance is also supported by data from deep sequencing of BCR repertoire showing
distinct gene segment usage biases in both VH and VL sequences within the naïve and
memory compartments in AChR-MG and MuSK-MG [101].

Figure 3. Schematic diagram of pathogenesis of AChR-MG. Impaired tolerance to the AChR is


the result of thymoma, thymic dysplasia or due to certain genetic background, which results in
presentation of AChR to the naïve T cells by thymic myoid cells or antigen-presenting cells. Among
the environmental factors, certain drugs are known to cause de novo MG through alterations of
immune homeostasis (drug-induced MG is extensively covered in another paper in this special
edition). A number of T cell and B cell subtypes and their cytokines play roles in perturbation of
immune homeostasis that results in production of ACR antibodies. HLA: Human Leucocyte antigen;
CTLA4: cytotoxic T-lymphocyte-associated protein 4; TNFRSF11A: tumor necrosis factor receptor
4 superfamily, member 11a; AIRE: autoimmune regulator; Th1: T helper 1; Th2: T helper 2; Tfh:
T follicular helper; Treg: regulatory T cell; Bregs: regulatory B cells; IL: interleukin; BAFF: B cell
activating factor.

5.3. Role of Thymus in MG


Autoreactive T cells also play an important role in the development of the disease
(Figure 3). The T cell selection process takes place in the thymus, mainly in the thymic
medulla, where they undergo negative selection for self-antigens [102]. Thymic epithe-
lial cells present self-antigens to developing T cells, either directly or through antigen-
presenting cells. If the developing T cells bind strongly to these antigens, they are removed
from the repertoire. Mechanisms for removal include (1) clonal deletion, (2) the induction
J. Clin. Med. 2021, 10, 2235 10 of 17

of anergy, or (3) clonal diversion/transformation into regulatory T cells (Tregs). The AChRs,
as well as other muscle proteins such as ryanodine and titin, are expressed by thymic myoid
and epithelial cells [103]. A key player in T cell autoimmunity is the autoimmune regulator
(AIRE) transcription factor, which induces tolerance over autoimmunity, by helping with
self-antigen expression in thymic cells. This transcription factor is modulated by estrogen,
which may help explain the early female predominance of the disease [104].
MG is uniquely associated with thymus hyperplasia and thymoma. The presence of
ectopic germinal centers is associated with early-onset AChR-MG, but not with MuSK-
associated MG. The T cell selection process may be impaired in thymic hyperplasia and
thymoma. The latter can have defective AIRE expression and can lack thymic medulla,
which is involved in the negative selection of T cells, therefore contributing to the release of
autoreactive CD4+ and CD8+ T cells [105]. In addition, the ensuing T cell subset imbalance
and cytokine dysregulation leads to the activation of B cells in germinal centers and
differentiation into autoreactive B cells and plasma cells [106]. T follicular helper cells
(Tfh), which are required for the generation of germinal centers, produce IL-21 and induce
immunoglobulin class switching [107]. Microarray analysis in thymic T cells from MG
patients revealed a Th1/Th17/Tfh signature [108].
Tfh in lymphoid organs are difficult to assess, so some investigators evaluate the
frequency of circulating Tfh (CXC5+ CD4 T cells in peripheral blood) instead. Ashida et al.
demonstrated an increase in circulating Tfh with elevated expression of inducible T cell
costimulator (ICOS) in a cohort of treatment naïve AChR-MG patients compared to con-
trols [109]. Interestingly, the frequency of circulating Tfh correlated with the severity of
MG, and its response to treatment in that study.

5.3.1. Role of T Cells and Cytokines in the Development of MG


Although MG is a B cell-mediated disease, CD4+ T cells and their cytokines contribute
to the development of disease. Animal studies showed that mice with depletion of CD4+ T
cells or class II major histocompatibility complex (MHC II) did not develop the experimental
autoimmune MG after sensitization to AChRs [110,111], further confirming their role.
Wu et al. demonstrated that the induction of tolerance to T cell epitopes prevented the
development of MG phenotype in a murine model after immunization with AChRs [112].
Patients with MG have autoreactive Th1 and Th2 cells, but their specific role in
autoantibody production is not clear. Cytokine measurements in sera by ELISA or by flow
cytometry often reveal conflicting results. Th2 cytokines such as IL-4 are known to play a
role in the induction of B cells; therefore, it is believed that a humoral Th2 response has a
direct role in the immunopathogenesis of the disease [113]. On the other hand, CD4+ T cells
with a Th1 profile have also been shown to play a role in MG. For example, patients with
MG have high levels of IFN-γ-secreting Th1 cells in the peripheral blood [114]. It is unclear
how IFN-γ promotes the production of autoantibodies, but it is believed to induce MHC II
and costimulatory molecules in adjacent tissues, such as myocytes, rendering them with
the ability to present antigens and promote antibody response. There is some evidence that
non-complement fixing isotypes such as anti-MuSK IgG4 are regulated by Th2 cytokines,
whereas complement fixing isotypes are regulated by Th1 cytokines [115,116]. Yet, a recent
study found an increase in Th2 cytokine IL-4, in addition to IL-10, IL-17 and IL-21 in CD4+
T cells in patients with AChR-MG, when compared to healthy controls [117].
Th17 T cells are associated with tissue-specific autoimmune inflammatory disorders
by activating immune cells and promoting their migration into tissues, thus enhancing
inflammation overall. These cells are postulated to play a key role in the pathogenesis of
MG. Multiple mechanisms are implicated, including the release of interleukin 17 (IL-17),
among other cytokines, which indirectly promote immunoglobulin production. Th17 cells
also affect the balance of the cytokine profile of Th1 and Th2 cells, thus influencing antibody
production [118]. Several studies have shown that patients with MG have elevated levels of
Th17 cells and IL-17, which correlate with disease severity and antibody titers [119,120]. In
the study by Cao and colleagues, autoreactive T cells from AChR-MG exhibited increased
J. Clin. Med. 2021, 10, 2235 11 of 17

IL-17, IFN-γ, and GM-CSF and diminished IL-10 production [121]. On the other hand, T
cell and cytokine profile may change during MG crisis. Huan et al. demonstrated that
patients with MG crisis had significantly elevated levels of Th17 as well as Th2-related
cytokines IL-4 and IL-13 compared to 6 months post-ventilatory support [122]. Higher
frequencies of Th1 and Th17 cytokines were also observed in MuSK-associated MG [123].
During infections, there is likely a perturbation in the cytokine milieu that increases the
risk of MG exacerbation.

5.3.2. Regulatory T Cells (Tregs), Regulatory B Cells (Bregs), and B Cell-Activating Factor
(BAFF) Signaling in MG
Autoimmune diseases such as MG are precipitated when there is an altered balance
between autoreactive T and B cells, and regulatory cell types that suppress them. The
latter include Tregs and regulatory B cells, both of which are phenotypically diverse. Tregs
suppress the function of other effector T cells and antigen-presenting cells by releasing
anti-inflammatory cytokines, such as IL-10 and transforming growth factor beta (TGF-ß),
and through the expression of forkhead box protein 3 (FoxP3), among other mechanisms.
Th17/Treg imbalance has been reported in patients with MG, especially in those with
generalized forms and thymomatous MG [108,124]. Impaired suppressive function of
CD4+ Tregs from thymus and peripheral blood cells of MG patients has been demonstrated,
though the number of CD4+ Tregs was unchanged in most studies [125–128].
Regarding Bregs, multiple subsets of B cells with overlapping markers have been
reported to produce IL-10 and suppress pro-inflammatory responses, but the establish-
ment of consensus in the field has been hampered by the lack of a unique transcription
factor [129]. We and other investigators have reported that MG patients exhibited a de-
crease in the frequency of CD19+ CD1dhi CD5+ and CD19+ CD24hi CD38hi Breg subsets and
IL-10-producing B cells within each subset [130–133]. Taken together, the expansion of
Tregs and Bregs or the restoration of its suppressive function is a potential therapeutic
strategy in MG.
In contrast to the impaired number and/or function of Tregs and Bregs, there is some
evidence of enhanced BAFF signaling in MG. BAFF is a member of the tumor necrosis factor
family. BAFF signaling through interaction with BAFF-receptor (BAFF-R) is essential for B
cell survival, maturation, and their development into plasmablasts and plasma cells [134].
The levels of circulating BAFF, which is secreted by myeloid cells, are increased in the sera of
MG patients [135,136]. In addition, MG patients exhibit an increase in BAFF-R+ B cells [137].
These findings support a role for dysregulated BAFF signaling in MG pathogenesis.

6. Conclusions
MG affects all age groups, with peaks in younger women and older men. There is a
great variability in the geographical/regional rates of MG, with the incidence and preva-
lence rates increasing overall, the latter partly due to better awareness and improvements in
the diagnosis of the disease. Juvenile MG is more common in people of Asian and African
descent. A subset of AChR-MG is caused by a thymoma or thymic hyperplasia. The rest of
AChR-MG as well as all of MuSK, Lrp4, and seronegative-MG cases are primarily autoim-
mune in nature, though influenced by genetic background and environmental factors that
are fully understood. Although primarily an antibody-mediated disorder, different T and
B cell subsets, including Th2, Th1, Th17, Tfh, Treg and Breg, and their related cytokines
play important roles in MG pathogenesis. A deeper understanding of MG subgroups and
their distinct immunopathogenic mechanisms will result in the identification of therapeutic
targets and the development of targeted treatment strategies.

Author Contributions: L.D. drafted the manuscript. R.W. contributed to drafting the section on
pediatric MG. K.R. and B.S. revised the manuscript. All authors have read and agreed to the published
version of the manuscript.
Funding: This research received no external funding.
J. Clin. Med. 2021, 10, 2235 12 of 17

Institutional Review Board Statement: Not applicable.


Informed Consent Statement: Not applicable.
Data Availability Statement: No new data were created or analyzed in this study. Data sharing is
not applicable to this article.
Conflicts of Interest: K.R. has received honoraria for consultations and serving on advisory boards
for Alexion, Kabafusion and Grifols. B.S. and K.R. have received funding from Alexion for conducting
clinical trials on MG and ALS. L.D. and R.W. declare no conflict of interest.

References
1. McGrogan, A.; Sneddon, S.; de Vries, C.S. The Incidence of Myasthenia Gravis: A Systematic Literature Review. Neuroepidemiology
2010, 34, 171–183. [CrossRef] [PubMed]
2. Patrick, J.; Lindstrom, J. Autoimmune Response to Acetylcholine Receptor. Science 1973, 180, 871–872. [CrossRef] [PubMed]
3. Fambrough, D.M.; Drachman, D.B.; Satyamurti, S. Neuromuscular Junction in Myasthenia Gravis: Decreased Acetylcholine
Receptors. Science 1973, 182, 293–295. [CrossRef] [PubMed]
4. McConville, J.; Farrugia, M.E.; Beeson, D.; Kishore, U.; Metcalfe, R.; Newsom-Davis, J.; Vincent, A. Detection and Characterization
of MuSK Antibodies in Seronegative Myasthenia Gravis. Ann. Neurol. 2004, 55, 580–584. [CrossRef]
5. Hoch, W.; McConville, J.; Helms, S.; Newsom-Davis, J.; Melms, A.; Vincent, A. Auto-Antibodies to the Receptor Tyrosine Kinase
MuSK in Patients with Myasthenia Gravis without Acetylcholine Receptor Antibodies. Nat. Med. 2001, 7, 365–368. [CrossRef]
6. Higuchi, O.; Hamuro, J.; Motomura, M.; Yamanashi, Y. Autoantibodies to Low-Density Lipoprotein Receptor-Related Protein 4 in
Myasthenia Gravis. Ann. Neurol. 2011, 69, 418–422. [CrossRef]
7. Pevzner, A.; Schoser, B.; Peters, K.; Cosma, N.-C.; Karakatsani, A.; Schalke, B.; Melms, A.; Kröger, S. Anti-LRP4 Autoantibodies in
AChR- and MuSK-Antibody-Negative Myasthenia Gravis. J. Neurol. 2012, 259, 427–435. [CrossRef]
8. Gasperi, C.; Melms, A.; Schoser, B.; Zhang, Y.; Meltoranta, J.; Risson, V.; Schaeffer, L.; Schalke, B.; Kröger, S. Anti-Agrin
Autoantibodies in Myasthenia Gravis. Neurology 2014, 82, 1976–1983. [CrossRef]
9. Zhang, B.; Shen, C.; Bealmear, B.; Ragheb, S.; Xiong, W.-C.; Lewis, R.A.; Lisak, R.P.; Mei, L. Autoantibodies to Agrin in Myasthenia
Gravis Patients. PLoS ONE 2014, 9, e91816. [CrossRef]
10. Szczudlik, P.; Szyluk, B.; Lipowska, M.; Ryniewicz, B.; Kubiszewska, J.; Dutkiewicz, M.; Gilhus, N.E.; Kostera-Pruszczyk, A.
Antititin Antibody in Early- and Late-Onset Myasthenia Gravis. Acta Neurol. Scand. 2014, 130, 229–233. [CrossRef]
11. Kufukihara, K.; Watanabe, Y.; Inagaki, T.; Takamatsu, K.; Nakane, S.; Nakahara, J.; Ando, Y.; Suzuki, S. Cytometric Cell-Based
Assays for Anti-Striational Antibodies in Myasthenia Gravis with Myositis and/or Myocarditis. Sci. Rep. 2019, 9, 5284. [CrossRef]
12. Grob, D.; Brunner, N.; Namba, T.; Pagala, M. Lifetime Course of Myasthenia Gravis. Muscle Nerve 2008, 37, 141–149. [CrossRef]
13. Phillips, L.H. The Epidemiology of Myasthenia Gravis. Ann. N. Y. Acad. Sci. 2003, 998, 407–412. [CrossRef]
14. Somnier, F.E.; Engel, P.J.H. The Occurrence of Anti-Titin Antibodies and Thymomas: A Population Survey of MG 1970–1999.
Neurology 2002, 59, 92–98. [CrossRef] [PubMed]
15. MacDonald, B.K.; Cockerell, O.C.; Sander, J.W.; Shorvon, S.D. The Incidence and Lifetime Prevalence of Neurological Disorders
in a Prospective Community-Based Study in the UK. Brain 2000, 123 (Pt 4), 665–676. [CrossRef]
16. Fang, W.; Li, Y.; Mo, R.; Wang, J.; Qiu, L.; Ou, C.; Lin, Z.; Huang, Z.; Feng, H.; He, X.; et al. Hospital and Healthcare
Insurance System Record–Based Epidemiological Study of Myasthenia Gravis in Southern and Northern China. Neurol. Sci. 2020,
41, 1211–1223. [CrossRef] [PubMed]
17. Lee, H.S.; Lee, H.S.; Shin, H.Y.; Choi, Y.C.; Kim, S.M. The Epidemiology of Myasthenia Gravis in Korea. Yonsei Med. J. 2016,
57, 419–425. [CrossRef] [PubMed]
18. Park, S.Y.; Lee, J.Y.; Lim, N.G.; Hong, Y.H. Incidence and Prevalence of Myasthenia Gravis in Korea: A Population-Based Study
Using the National Health Insurance Claims Database. J. Clin. Neurol. 2016, 12, 340–344. [CrossRef]
19. Bettini, M.; Chaves, M.; Cristiano, E.; Pagotto, V.; Perez, L.; Giunta, D.; Rugiero, M. Incidence of Autoimmune Myasthenia Gravis
in a Health Maintenance Organization in Buenos Aires, Argentina. Neuroepidemiology 2017, 48, 119–123. [CrossRef]
20. Carr, A.S.; Cardwell, C.R.; McCarron, P.O.; McConville, J. A Systematic Review of Population Based Epidemiological Studies in
Myasthenia Gravis. Bmc Neurol. 2010, 10, 46. [CrossRef]
21. Ciafaloni, E. Myasthenia Gravis and Congenital Myasthenic Syndromes. Contin. Lifelong Learn. Neurol. 2019, 25, 1767–1784.
[CrossRef]
22. Oh, S.J.; Morgan, M.B.; Lu, L.; Hatanaka, Y.; Hemmi, S.; Young, A.; Claussen, G.C. Racial Differences in Myasthenia Gravis in
Alabama. Muscle Nerve 2009, 39, 328–332. [CrossRef]
23. Phillips, L.H.; Torner, J.C.; Anderson, M.S.; Cox, G.M. The Epidemiology of Myasthenia Gravis in Central and Western Virginia.
Neurology 1992, 42, 1888–1893. [CrossRef] [PubMed]
24. Alshekhlee, A.; Miles, J.D.; Katirji, B.; Preston, D.C.; Kaminski, H.J. Incidence and Mortality Rates of Myasthenia Gravis and
Myasthenic Crisis in US Hospitals. Neurology 2009, 72, 1548–1554. [CrossRef]
25. Heckmann, J.M.; Owen, E.P.; Little, F. Myasthenia Gravis in South Africans: Racial Differences in Clinical Manifestations.
Neuromuscul. Disord. 2007, 17, 929–934. [CrossRef] [PubMed]
J. Clin. Med. 2021, 10, 2235 13 of 17

26. Peragallo, J.H.; Bitrian, E.; Kupersmith, M.J.; Zimprich, F.; Whittaker, T.J.; Lee, M.S.; Bruce, B.B. Relationship between Age,
Gender, and Race in Patients Presenting with Myasthenia Gravis with Only Ocular Manifestations. J. Neuroophthalmol. 2016,
36, 29–32. [CrossRef]
27. Boldingh, M.I.; Maniaol, A.; Brunborg, C.; Dekker, L.; Lipka, A.; Niks, E.H.; Verschuuren, J.; Tallaksen, C. Prevalence and Clinical
Aspects of Immigrants with Myasthenia Gravis in Northern Europe. Muscle Nerve 2017, 55, 819–827. [CrossRef] [PubMed]
28. Abukhalil, F.; Mehta, B.; Saito, E.; Mehta, S.; McMurtray, A. Gender and Ethnicity Based Differences in Clinical and Laboratory
Features of Myasthenia Gravis. Autoimmune Dis. 2015, 2015, 197893. [CrossRef]
29. Deymeer, F.; Gungor-Tuncer, O.; Yilmaz, V.; Parman, Y.; Serdaroglu, P.; Ozdemir, C.; Vincent, A.; Saruhan-Direskeneli, G. Clinical
Comparison of Anti-MuSK- vs Anti-AChR-Positive and Seronegative Myasthenia Gravis. Neurology 2007, 68, 609–611. [CrossRef]
30. Albuquerque, E.X.; Pereira, E.F.R.; Alkondon, M.; Rogers, S.W. Mammalian Nicotinic Acetylcholine Receptors: From Structure to
Function. Physiol. Rev. 2009, 89, 73–120. [CrossRef]
31. Tzartos, S.J.; Barkas, T.; Cung, M.T.; Mamalaki, A.; Marraud, M.; Orlewski, P.; Papanastasiou, D.; Sakarellos, C.; Sakarellos-
Daitsiotis, M.; Tsantili, P.; et al. Anatomy of the Antigenic Structure of a Large Membrane Autoantigen, the Muscle-Type Nicotinic
Acetylcholine Receptor. Immunol. Rev. 1998, 163, 89–120. [CrossRef]
32. Kordas, G.; Lagoumintzis, G.; Sideris, S.; Poulas, K.; Tzartos, S.J. Direct Proof of the In Vivo Pathogenic Role of the AChR
Autoantibodies from Myasthenia Gravis Patients. PLoS ONE 2014, 9, e108327. [CrossRef]
33. Lennon, V.A.; McCormick, D.J.; Lambert, E.H.; Griesmann, G.E.; Atassi, M.Z. Region of Peptide 125-147 of Acetylcholine Receptor
Alpha Subunit Is Exposed at Neuromuscular Junction and Induces Experimental Autoimmune Myasthenia Gravis, T-Cell
Immunity, and Modulating Autoantibodies. Proc. Natl. Acad. Sci. USA 1985, 82, 8805–8809. [CrossRef]
34. Fostieri, E.; Beeson, D.; Tzartos, S.J. The Conformation of the Main Immunogenic Region on the α-Subunit of Muscle Acetylcholine
Receptor Is Affected by Neighboring Receptor Subunits. Febs Lett. 2000, 481, 127–130. [CrossRef]
35. Morgan, B.P.; Chamberlain-Banoub, J.; Neal, J.W.; Song, W.; Mizuno, M.; Harris, C.L. The Membrane Attack Pathway of
Complement Drives Pathology in Passively Induced Experimental Autoimmune Myasthenia Gravis in Mice. Clin. Exp. Immunol.
2006, 146, 294–302. [CrossRef]
36. Rødgaard, A.; Nielsen, F.C.; Djurup, R.; Somnier, F.; Gammeltoft, S. Acetylcholine Receptor Antibody in Myasthenia Gravis:
Predominance of IgG Subclasses 1 and 3. Clin. Exp. Immunol. 1987, 67, 82–88.
37. Drachman, D.B.; Angus, C.W.; Adams, R.N.; Michelson, J.D.; Hoffman, G.J. Myasthenic Antibodies Cross-Link Acetylcholine
Receptors to Accelerate Degradation. N. Engl. J. Med. 1978, 298, 1116–1122. [CrossRef]
38. Hara, H.; Hayashi, K.; Ohta, K.; Itoh, N.; Nishitani, H.; Ohta, M. Detection and Characterization of Blocking-Type Anti-
Acetylcholine Receptor Antibodies in Sera from Patients with Myasthenia Gravis. Clin. Chem. 1993, 39, 2053–2057. [CrossRef]
39. Wintzen, A.R.; Plomp, J.J.; Molenaar, P.C.; van Dijk, J.G.; van Kempen, G.T.; Vos, R.M.; Wokke, J.H.; Vincent, A. Acquired
Slow-Channel Syndrome: A Form of Myasthenia Gravis with Prolonged Open Time of the Acetylcholine Receptor Channel. Ann.
Neurol. 1998, 44, 657–664. [CrossRef]
40. Howard, F.M.; Lennon, V.A.; Finley, J.; Matsumoto, J.; Elveback, L.R. Clinical Correlations of Antibodies That Bind, Block, or
Modulate Human Acetylcholine Receptors in Myasthenia Gravis. Ann. N. Y. Acad. Sci. 1987, 505, 526–538. [CrossRef]
41. Gomez, C.M.; Richman, D.P. Anti-Acetylcholine Receptor Antibodies Directed against the Alpha-Bungarotoxin Binding Site
Induce a Unique Form of Experimental Myasthenia. Proc. Natl. Acad. Sci. USA 1983, 80, 4089–4093. [CrossRef]
42. Oosterhuis, H.J. The Natural Course of Myasthenia Gravis: A Long Term Follow up Study. J. Neurol. Neurosurg. Psychiatry 1989,
52, 1121–1127. [CrossRef]
43. Hong, Y.-H.; Kwon, S.-B.; Kim, B.-J.; Kim, B.J.; Kim, S.H.; Kim, J.K.; Park, K.-S.; Park, K.-J.; Sung, J.-J.; Sohn, E.H.; et al. Prognosis
of Ocular Myasthenia in Korea: A Retrospective Multicenter Analysis of 202 Patients. J. Neurol. Sci. 2008, 273, 10–14. [CrossRef]
44. Hendricks, T.M.; Bhatti, M.T.; Hodge, D.O.; Chen, J.J. Incidence, Epidemiology, and Transformation of Ocular Myasthenia Gravis:
A Population-Based Study. Am. J. Ophthalmol. 2019, 205, 99–105. [CrossRef]
45. Gilhus, N.E. Myasthenia Gravis. N. Engl. J. Med. 2016, 375, 2570–2581. [CrossRef]
46. Sih, M.; Soliven, B.; Mathenia, N.; Jacobsen, J.; Rezania, K. Head-Drop: A Frequent Feature of Late-Onset Myasthenia Gravis:
Head-Drop in Myasthenia Gravis. Muscle Nerve 2017, 56, 441–444. [CrossRef]
47. Grob, D. Course and Management of Myasthenia Gravis. J. Am. Med. Assoc. 1953, 153, 529–532. [CrossRef]
48. Kaminski, H.J.; Maas, E.; Spiegel, P.; Ruff, R.L. Why Are Eye Muscles Frequently Involved in Myasthenia Gravis? Neurology 1990,
40, 1663–1669. [CrossRef]
49. MacLennan, C.; Beeson, D.; Buijs, A.-M.; Vincent, A.; Newsom-Davis, J. Acetylcholine Receptor Expression in Human Extraocular
Muscles and Their Susceptibility to Myasthenia Gravis. Ann. Neurol. 1997, 41, 423–431. [CrossRef]
50. Narayanaswami, P.; Sanders, D.B.; Wolfe, G.; Benatar, M.; Cea, G.; Evoli, A.; Gilhus, N.E.; Illa, I.; Kuntz, N.L.; Massey, J.; et al.
International Consensus Guidance for Management of Myasthenia Gravis: 2020 Update. Neurology 2021, 96, 114–122. [CrossRef]
[PubMed]
51. Gilhus, N.E.; Verschuuren, J.J. Myasthenia Gravis: Subgroup Classification and Therapeutic Strategies. Lancet Neurol. 2015, 14,
1023–1036. [CrossRef]
52. Vandiedonck, C.; Beaurain, G.; Giraud, M.; Hue-Beauvais, C.; Eymard, B.; Tranchant, C.; Gajdos, P.; Dausset, J.; Garchon, H.-J.
Pleiotropic Effects of the 8.1 HLA Haplotype in Patients with Autoimmune Myasthenia Gravis and Thymus Hyperplasia. Proc.
Natl. Acad. Sci. USA 2004, 101, 15464–15469. [CrossRef] [PubMed]
J. Clin. Med. 2021, 10, 2235 14 of 17

53. Spagni, G.; Todi, L.; Monte, G.; Valentini, M.; Di Sante, G.; Damato, V.; Marino, M.; Evoli, A.; Lantieri, F.; Provenzano, C. Human
Leukocyte Antigen Class II Associations in Late-onset Myasthenia Gravis. Ann. Clin. Transl. Neurol. 2021, 8, 656–665. [CrossRef]
[PubMed]
54. Saruhan-Direskeneli, G.; Hughes, T.; Yilmaz, V.; Durmus, H.; Adler, A.; Alahgholi-Hajibehzad, M.; Aysal, F.; Yentür, S.P.; Akalin,
M.A.; Dogan, O.; et al. Genetic Heterogeneity within the HLA Region in Three Distinct Clinical Subgroups of Myasthenia Gravis.
Clin. Immunol. 2016, 166–167, 81–88. [CrossRef] [PubMed]
55. Maniaol, A.H.; Elsais, A.; Lorentzen, Å.R.; Owe, J.F.; Viken, M.K.; Sæther, H.; Flåm, S.T.; Bråthen, G.; Kampman, M.T.; Midgard,
R.; et al. Late Onset Myasthenia Gravis Is Associated with HLA DRB1*15:01 in the Norwegian Population. PLoS ONE 2012,
7, e36603. [CrossRef]
56. Renton, A.E.; Pliner, H.A.; Provenzano, C.; Evoli, A.; Ricciardi, R.; Nalls, M.A.; Marangi, G.; Abramzon, Y.; Arepalli, S.; Chong, S.;
et al. A Genome-Wide Association Study of Myasthenia Gravis. Jama Neurol. 2015, 72, 396–404. [CrossRef]
57. Gilhus, N.E.; Skeie, G.O.; Romi, F.; Lazaridis, K.; Zisimopoulou, P.; Tzartos, S. Myasthenia Gravis—Autoantibody Characteristics
and Their Implications for Therapy. Nat Rev Neurol. 2016, 12, 259–268. [CrossRef]
58. Bernard, C.; Frih, H.; Pasquet, F.; Kerever, S.; Jamilloux, Y.; Tronc, F.; Guibert, B.; Isaac, S.; Devouassoux, M.; Chalabreysse, L.; et al.
Thymoma Associated with Autoimmune Diseases: 85 Cases and Literature Review. Autoimmun. Rev. 2016, 15, 82–92. [CrossRef]
59. Muñiz-Castrillo, S.; Vogrig, A.; Honnorat, J. Associations between HLA and Autoimmune Neurological Diseases with Autoanti-
bodies. Auto Immun. Highlights 2020, 11, 2. [CrossRef]
60. Kim, N.; Stiegler, A.L.; Cameron, T.O.; Hallock, P.T.; Gomez, A.M.; Huang, J.H.; Hubbard, S.R.; Dustin, M.L.; Burden, S.J. Lrp4 Is
a Receptor for Agrin and Forms a Complex with MuSK. Cell 2008, 135, 334–342. [CrossRef]
61. Guptill, J.T.; Sanders, D.B.; Evoli, A. Anti-Musk Antibody Myasthenia Gravis: Clinical Findings and Response to Treatment in
Two Large Cohorts. Muscle Nerve 2011, 44, 36–40. [CrossRef] [PubMed]
62. Cole, R.N.; Reddel, S.W.; Gervásio, O.L.; Phillips, W.D. Anti-MuSK Patient Antibodies Disrupt the Mouse Neuromuscular
Junction. Ann. Neurol. 2008, 63, 782–789. [CrossRef]
63. Niks, E.H.; van Leeuwen, Y.; Leite, M.I.; Dekker, F.W.; Wintzen, A.R.; Wirtz, P.W.; Vincent, A.; van Tol, M.J.D.; Jol-van der Zijde,
C.M.; Verschuuren, J.J.G.M. Clinical Fluctuations in MuSK Myasthenia Gravis Are Related to Antigen-Specific IgG4 Instead of
IgG1. J. Neuroimmunol. 2008, 195, 151–156. [CrossRef] [PubMed]
64. Poulas, K.; Koutsouraki, E.; Kordas, G.; Kokla, A.; Tzartos, S.J. Anti-MuSK- and Anti-AChR-Positive Myasthenia Gravis Induced
by d-Penicillamine. J. Neuroimmunol. 2012, 250, 94–98. [CrossRef]
65. Plomp, J.J.; Huijbers, M.G.; van der Maarel, S.M.; Verschuuren, J.J. Pathogenic IgG4 Subclass Autoantibodies in MuSK Myasthenia
Gravis: MuSK Myasthenia Gravis IgG4. Ann. N. Y. Acad. Sci. 2012, 1275, 114–122. [CrossRef] [PubMed]
66. Okada, K.; Inoue, A.; Okada, M.; Murata, Y.; Kakuta, S.; Jigami, T.; Kubo, S.; Shiraishi, H.; Eguchi, K.; Motomura, M.; et al. The
Muscle Protein Dok-7 Is Essential for Neuromuscular Synaptogenesis. Science 2006, 312, 1802–1805. [CrossRef]
67. Huijbers, M.G.; Zhang, W.; Klooster, R.; Niks, E.H.; Friese, M.B.; Straasheijm, K.R.; Thijssen, P.E.; Vrolijk, H.; Plomp, J.J.; Vogels, P.;
et al. MuSK IgG4 Autoantibodies Cause Myasthenia Gravis by Inhibiting Binding between MuSK and Lrp4. Proc. Natl. Acad. Sci.
USA 2013, 110, 20783–20788. [CrossRef] [PubMed]
68. Koneczny, I.; Cossins, J.; Waters, P.; Beeson, D.; Vincent, A. MuSK Myasthenia Gravis IgG4 Disrupts the Interaction of LRP4
with MuSK but Both IgG4 and IgG1-3 Can Disperse Preformed Agrin-Independent AChR Clusters. PLoS ONE 2013, 8, e80695.
[CrossRef]
69. Sanders, D.B.; El-Salem, K.; Massey, J.M.; McConville, J.; Vincent, A. Clinical Aspects of MuSK Antibody Positive Seronegative
MG. Neurology 2003, 60, 1978–1980. [CrossRef]
70. Pasnoor, M.; Wolfe, G.I.; Nations, S.; Trivedi, J.; Barohn, R.J.; Herbelin, L.; McVey, A.; Dimachkie, M.; Kissel, J.; Walsh, R.; et al.
Clinical Findings in MuSK-Antibody Positive Myasthenia Gravis: A U.S. Experience. Muscle Nerve 2010, 41, 370–374. [CrossRef]
71. Bartoccioni, E.; Scuderi, F.; Augugliaro, A.; Chiatamone Ranieri, S.; Sauchelli, D.; Alboino, P.; Marino, M.; Evoli, A. HLA Class II
Allele Analysis in MuSK-Positive Myasthenia Gravis Suggests a Role for DQ5. Neurology 2009, 72, 195–197. [CrossRef]
72. Alahgholi-Hajibehzad, M.; Yilmaz, V.; Gülsen-Parman, Y.; Aysal, F.; Oflazer, P.; Deymeer, F.; Saruhan-Direskeneli, G. Association
of HLA-DRB1∗14, -DRB1∗16 and -DQB1∗05 with MuSK-Myasthenia Gravis in Patients from Turkey. Hum. Immunol. 2013,
74, 1633–1635. [CrossRef] [PubMed]
73. Zhang, J.; Chen, Y.; Chen, J.; Huang, X.; Wang, H.; Li, Y.; Liu, W.; Feng, H. AChRAb and MuSKAb Double-Seropositive Myasthenia
Gravis: A Distinct Subtype? Neurol. Sci. 2021, 42, 863–869. [CrossRef]
74. Cortés-Vicente, E.; Gallardo, E.; Martínez, M.Á.; Díaz-Manera, J.; Querol, L.; Rojas-García, R.; Illa, I. Clinical Characteristics of
Patients With Double-Seronegative Myasthenia Gravis and Antibodies to Cortactin. Jama Neurol. 2016, 73, 1099. [CrossRef]
[PubMed]
75. Rivner, M.H.; Quarles, B.M.; Pan, J.; Yu, Z.; Howard, J.F.; Corse, A.; Dimachkie, M.M.; Jackson, C.; Vu, T.; Small, G.; et al. Clinical
Features of LRP4 /Agrin-antibody–Positive Myasthenia Gravis: A Multicenter Study. Muscle Nerve 2020, 62, 333–343. [CrossRef]
76. Zhang, B.; Tzartos, J.S.; Belimezi, M.; Ragheb, S.; Bealmear, B.; Lewis, R.A.; Xiong, W.-C.; Lisak, R.P.; Tzartos, S.J.; Mei, L.
Autoantibodies to Lipoprotein-Related Protein 4 in Patients with Double-Seronegative Myasthenia Gravis. Arch. Neurol. 2012,
69, 445–451. [CrossRef]
77. Rivner, M.H.; Liu, S.; Quarles, B.; Fleenor, B.; Shen, C.; Pan, J.; Mei, L. Agrin and Low-Density Lipoprotein-Related Receptor
Protein 4 Antibodies in Amyotrophic Lateral Sclerosis Patients. Muscle Nerve 2017, 55, 430–432. [CrossRef]
J. Clin. Med. 2021, 10, 2235 15 of 17

78. Tzartos, J.S.; Zisimopoulou, P.; Rentzos, M.; Karandreas, N.; Zouvelou, V.; Evangelakou, P.; Tsonis, A.; Thomaidis, T.; Lauria, G.;
Andreetta, F.; et al. LRP4 Antibodies in Serum and CSF from Amyotrophic Lateral Sclerosis Patients. Ann. Clin. Transl. Neurol.
2014, 1, 80–87. [CrossRef] [PubMed]
79. Lai, C.-H.; Tseng, H.-F. Nationwide Population-Based Epidemiological Study of Myasthenia Gravis in Taiwan. Neuroepidemiology
2010, 35, 66–71. [CrossRef]
80. Chiu, H.C.; Vincent, A.; Newsom-Davis, J.; Hsieh, K.H.; Hung, T. Myasthenia Gravis: Population Differences in Disease
Expression and Acetylcholine Receptor Antibody Titers between Chinese and Caucasians. Neurology 1987, 37, 1854–1857.
[CrossRef]
81. Chiang, L.M.; Darras, B.T.; Kang, P.B. Juvenile Myasthenia Gravis. Muscle Nerve 2009, 39, 423–431. [CrossRef]
82. Castro, D.; Derisavifard, S.; Anderson, M.; Greene, M.; Iannaccone, S. Juvenile Myasthenia Gravis: A Twenty-Year Experience. J.
Clin. Neuromuscul. Dis. 2013, 14, 95–102. [CrossRef]
83. Barraud, C.; Desguerre, I.; Barnerias, C.; Gitiaux, C.; Boulay, C.; Chabrol, B. Clinical Features and Evolution of Juvenile Myasthenia
Gravis in a French Cohort. Muscle Nerve 2018, 57, 603–609. [CrossRef] [PubMed]
84. Parr, J.R.; Jayawant, S. Childhood Myasthenia: Clinical Subtypes and Practical Management. Dev. Med. Child. Neurol. 2007,
49, 629–635. [CrossRef] [PubMed]
85. Lindner, A.; Schalke, B.; Toyka, K.V. Outcome in Juvenile-Onset Myasthenia Gravis: A Retrospective Study with Long-Term
Follow-up of 79 Patients. J. Neurol. 1997, 244, 515–520. [CrossRef] [PubMed]
86. Vanikieti, K.; Lowwongngam, K.; Padungkiatsagul, T.; Visudtibhan, A.; Poonyathalang, A. Juvenile Ocular Myasthenia Gravis:
Presentation and Outcome of a Large Cohort. Pediatr. Neurol. 2018, 87, 36–41. [CrossRef] [PubMed]
87. Gui, M.; Luo, X.; Lin, J.; Li, Y.; Zhang, M.; Zhang, X.; Yang, M.; Wang, W.; Bu, B. Long-Term Outcome of 424 Childhood-Onset
Myasthenia Gravis Patients. J. Neurol. 2015, 262, 823–830. [CrossRef]
88. Pineles, S.L.; Avery, R.A.; Moss, H.E.; Finkel, R.; Blinman, T.; Kaiser, L.; Liu, G.T. Visual and Systemic Outcomes in Pediatric
Ocular Myasthenia Gravis. Am. J. Ophthalmol. 2010, 150, 453–459.e3. [CrossRef] [PubMed]
89. Anlar, B.; Senbil, N.; Köse, G.; Değerliyurt, A. Serological Follow-up in Juvenile Myasthenia: Clinical and Acetylcholine Receptor
Antibody Status of Patients Followed for at Least 2 Years. Neuromuscul. Disord. 2005, 15, 355–357. [CrossRef]
90. Skjei, K.L.; Lennon, V.A.; Kuntz, N.L. Muscle Specific Kinase Autoimmune Myasthenia Gravis in Children: A Case Series.
Neuromuscul. Disord. 2013, 23, 874–882. [CrossRef]
91. Zisimopoulou, P.; Evangelakou, P.; Tzartos, J.; Lazaridis, K.; Zouvelou, V.; Mantegazza, R.; Antozzi, C.; Andreetta, F.; Evoli, A.;
Deymeer, F.; et al. A Comprehensive Analysis of the Epidemiology and Clinical Characteristics of Anti-LRP4 in Myasthenia
Gravis. J. Autoimmun. 2014, 52, 139–145. [CrossRef]
92. Li, M.; Han, J.; Zhang, Y.; Lv, J.; Zhang, J.; Zhao, X.; Ren, L.; Fang, H.; Yang, J.; Zhang, Y.; et al. Clinical Analysis of Chinese
Anti-Low-Density-Lipoprotein-Receptor-Associated Protein 4 Antibodies in Patients with Myasthenia Gravis. Eur. J. Neurol. 2019,
26, 1296-e84. [CrossRef] [PubMed]
93. Namba, T.; Brown, S.B.; Grob, D. Neonatal Myasthenia Gravis: Report of Two Cases and Review of the Literature. Pediatrics 1970,
45, 488–504. [PubMed]
94. Vernet-der Garabedian, B.; Lacokova, M.; Eymard, B.; Morel, E.; Faltin, M.; Zajac, J.; Sadovsky, O.; Dommergues, M.; Tripon, P.;
Bach, J.F. Association of Neonatal Myasthenia Gravis with Antibodies against the Fetal Acetylcholine Receptor. J. Clin. Investig.
1994, 94, 555–559. [CrossRef] [PubMed]
95. Hacohen, Y.; Jacobson, L.W.; Byrne, S.; Norwood, F.; Lall, A.; Robb, S.; Dilena, R.; Fumagalli, M.; Born, A.P.; Clarke, D.; et al. Fetal
Acetylcholine Receptor Inactivation Syndrome: A Myopathy Due to Maternal Antibodies. Neurol. Neuroimmunol. Neuroinflamm.
2015, 2, e57. [CrossRef]
96. Parr, J.R.; Andrew, M.J.; Finnis, M.; Beeson, D.; Vincent, A.; Jayawant, S. How Common Is Childhood Myasthenia? The UK
Incidence and Prevalence of Autoimmune and Congenital Myasthenia. Arch. Dis. Child. 2014, 99, 539–542. [CrossRef]
97. Koneczny, I.; Herbst, R. Myasthenia Gravis: Pathogenic Effects of Autoantibodies on Neuromuscular Architecture. Cells 2019,
8, 671. [CrossRef]
98. Ruff, R.L.; Lisak, R.P. Nature and Action of Antibodies in Myasthenia Gravis. Neurol. Clin. 2018, 36, 275–291. [CrossRef]
99. Ruff, R.L.; Lennon, V.A. How Myasthenia Gravis Alters the Safety Factor for Neuromuscular Transmission. J. Neuroimmunol.
2008, 201–202, 13–20. [CrossRef]
100. Lee, J.; Stathopoulos, P.; Gupta, S.; Bannock, J.M.; Barohn, R.J.; Cotzomi, E.; Dimachkie, M.M.; Jacobson, L.; Lee, C.S.; Morbach, H.;
et al. Compromised Fidelity of B-cell Tolerance Checkpoints in AChR and MuSK Myasthenia Gravis. Ann. Clin. Transl. Neurol.
2016, 3, 443–454. [CrossRef]
101. Vander Heiden, J.A.; Stathopoulos, P.; Zhou, J.Q.; Chen, L.; Gilbert, T.J.; Bolen, C.R.; Barohn, R.J.; Dimachkie, M.M.; Ciafaloni,
E.; Broering, T.J.; et al. Dysregulation of B Cell Repertoire Formation in Myasthenia Gravis Patients Revealed through Deep
Sequencing. J. Immunol. 2017, 198, 1460–1473. [CrossRef] [PubMed]
102. Takaba, H.; Takayanagi, H. The Mechanisms of T Cell Selection in the Thymus. Trends Immunol. 2017, 38, 805–816. [CrossRef]
[PubMed]
103. Poëa-Guyon, S.; Christadoss, P.; Le Panse, R.; Guyon, T.; De Baets, M.; Wakkach, A.; Bidault, J.; Tzartos, S.; Berrih-Aknin, S.
Effects of Cytokines on Acetylcholine Receptor Expression: Implications for Myasthenia Gravis. J. Immunol. 2005, 174, 5941–5949.
[CrossRef] [PubMed]
J. Clin. Med. 2021, 10, 2235 16 of 17

104. Dragin, N.; Bismuth, J.; Cizeron-Clairac, G.; Biferi, M.G.; Berthault, C.; Serraf, A.; Nottin, R.; Klatzmann, D.; Cumano, A.; Barkats,
M.; et al. Estrogen-Mediated Downregulation of AIRE Influences Sexual Dimorphism in Autoimmune Diseases. J. Clin. Investig.
2016, 126, 1525–1537. [CrossRef] [PubMed]
105. Truffault, F.; de Montpreville, V.; Eymard, B.; Sharshar, T.; Le Panse, R.; Berrih-Aknin, S. Thymic Germinal Centers and
Corticosteroids in Myasthenia Gravis: An Immunopathological Study in 1035 Cases and a Critical Review. Clin. Rev. Allerg.
Immunol. 2017, 52, 108–124. [CrossRef] [PubMed]
106. Vrolix, K.; Fraussen, J.; Losen, M.; Stevens, J.; Lazaridis, K.; Molenaar, P.C.; Somers, V.; Bracho, M.A.; Le Panse, R.; Stinissen,
P.; et al. Clonal Heterogeneity of Thymic B Cells from Early-Onset Myasthenia Gravis Patients with Antibodies against the
Acetylcholine Receptor. J. Autoimmun. 2014, 52, 101–112. [CrossRef]
107. Vinuesa, C.G.; Linterman, M.A.; Yu, D.; MacLennan, I.C.M. Follicular Helper T Cells. Annu. Rev. Immunol. 2016, 34, 335–368.
[CrossRef]
108. Gradolatto, A.; Nazzal, D.; Truffault, F.; Bismuth, J.; Fadel, E.; Foti, M.; Berrih-Aknin, S. Both Treg Cells and Tconv Cells Are
Defective in the Myasthenia Gravis Thymus: Roles of IL-17 and TNF-α. J. Autoimmun. 2014, 52, 53–63. [CrossRef]
109. Ashida, S.; Ochi, H.; Hamatani, M.; Fujii, C.; Kimura, K.; Okada, Y.; Hashi, Y.; Kawamura, K.; Ueno, H.; Takahashi, R.; et al.
Immune Skew of Circulating Follicular Helper T Cells Associates With Myasthenia Gravis Severity. Neurol. Neuroimmunol.
Neuroinflamm. 2021, 8. [CrossRef]
110. Kaul, R.; Shenoy, M.; Goluszko, E.; Christadoss, P. Major histocompatibility complex class II gene disruption prevents experimental
autoimmune myasthenia gravis. J. Immunol. 1994, 152, 3152–3157.
111. Christadoss, P.; Dauphinee, M.J. Immunotherapy for myasthenia gravis: A murine model. J. Immunol. 1986, 136, 2437–2440.
[PubMed]
112. Wu, B.; Deng, C.; Goluszko, E.; Christadoss, P. Tolerance to a dominant T cell epitope in the acetylcholine receptor molecule
induces epitope spread and suppresses murine myasthenia gravis. J. Immunol. 1997, 159, 3016–3023. [PubMed]
113. Balasa, B.; Sarvetnick, N. Is pathogenic humoral autoimmunity a Th1 response? Lessons from (for) myasthenia gravis. Immunol.
Today 2000, 21, 19–23. [CrossRef]
114. Link, J.; Söderström, M.; Ljungdahl, Å.; Höjeberg, B.; Olsson, T.; Xu, Z.; Fredrikson, S.; Wang, Z.-Y.; Link, H. Organ-specific
autoantigens induce interferon-γ and interleukin-4 mRNA expression in mononuclear cells in multiple sclerosis and myasthenia
gravis. Neurology 1994, 44, 728. [CrossRef] [PubMed]
115. Tanaka, A.; Moriyama, M.; Nakashima, H.; Miyake, K.; Hayashida, J.-N.; Maehara, T.; Shinozaki, S.; Kubo, Y.; Nakamura, S. Th2
and regulatory immune reactions contribute to IgG4 production and the initiation of Mikulicz disease. Arthritis Rheum. 2011,
64, 254–263. [CrossRef] [PubMed]
116. Stevens, T.L.; Bossie, A.; Sanders, V.M.; Fernandez-Botran, R.; Coffman, R.L.; Mosmann, T.R.; Vitetta, E.S. Regulation of antibody
isotype secretion by subsets of antigen-specific helper T cells. Nat. Cell Biol. 1988, 334, 255–258. [CrossRef]
117. Çebi, M.; Durmus, H.; Aysal, F.; Özkan, B.; Gül, G.E.; Çakar, A.; Hocaoglu, M.; Mercan, M.; Yentür, S.P.; Tütüncü, M.; et al. CD4+
T Cells of Myasthenia Gravis Patients Are Characterized by Increased IL-21, IL-4, and IL-17A Productions and Higher Presence
of PD-1 and ICOS. Front. Immunol. 2020, 11, 809. [CrossRef]
118. Uzawa, A.; Kuwabara, S.; Suzuki, S.; Imai, T.; Murai, H.; Ozawa, Y.; Yasuda, M.; Nagane, Y.; Utsugisawa, K. Roles of cytokines
and T cells in the pathogenesis of myasthenia gravis. Clin. Exp. Immunol. 2021, 203, 366–374. [CrossRef]
119. Xie, Y.; Li, H.-F.; Jiang, B.; Li, Y.; Kaminski, H.J.; Kusner, L.L. Elevated plasma interleukin-17A in a subgroup of Myasthenia
Gravis patients. Cytokine 2016, 78, 44–46. [CrossRef]
120. Wang, Z.; Wang, W.; Chen, Y.; Wei, D. T Helper Type 17 Cells Expand in Patients with Myasthenia-Associated Thymoma. Scand. J.
Immunol. 2012, 76, 54–61. [CrossRef]
121. Cao, Y.; Amezquita, R.A.; Kleinstein, S.H.; Stathopoulos, P.; Nowak, R.J.; O’Connor, K.C. Autoreactive T Cells from Patients with
Myasthenia Gravis Are Characterized by Elevated IL-17, IFN-γ, and GM-CSF and Diminished IL-10 Production. J. Immunol. 2016,
196, 2075–2084. [CrossRef]
122. Huan, X.; Luo, S.; Zhong, H.; Zheng, X.; Song, J.; Zhou, L.; Lu, J.; Wang, Y.; Xu, Y.; Xi, J.; et al. In-depth peripheral CD4 + T profile
correlates with myasthenic crisis. Ann. Clin. Transl. Neurol. 2021, 8, 749–762. [CrossRef] [PubMed]
123. Yi, J.; Guidon, A.; Sparks, S.; Osborne, R.; Juel, V.; Massey, J.; Sanders, D.; Weinhold, K.; Guptill, J. Characterization of CD4 and
CD8 T cell responses in MuSK myasthenia gravis. J. Autoimmun. 2014, 52, 130–138. [CrossRef] [PubMed]
124. Villegas, J.A.; Van Wassenhove, J.; Le Panse, R.; Berrih-Aknin, S.; Dragin, N. An imbalance between regulatory T cells and T
helper 17 cells in acetylcholine receptor-positive myasthenia gravis patients. Ann. N. Y. Acad. Sci. 2018, 1413, 154–162. [CrossRef]
[PubMed]
125. Balandina, A.; Lécart, S.; Dartevelle, P.; Saoudi, A.; Berrih-Aknin, S. Functional defect of regulatory CD4+CD25+ T cells in the
thymus of patients with autoimmune myasthenia gravis. Blood 2005, 105, 735–741. [CrossRef]
126. Battaglia, M.; Stabilini, A.; Roncarolo, M.-G. Rapamycin selectively expands CD4+CD25+FoxP3+ regulatory T cells. Blood 2005,
105, 4743–4748. [CrossRef]
127. Matsui, N.; Nakane, S.; Saito, F.; Ohigashi, I.; Nakagawa, Y.; Kurobe, H.; Takizawa, H.; Mitsui, T.; Kondo, K.; Kitagawa, T.; et al.
Undiminished regulatory T cells in the thymus of patients with myasthenia gravis. Neurology 2010, 74, 816–820. [CrossRef]
128. Thiruppathi, M.; Rowin, J.; Ganesh, B.; Sheng, J.R.; Prabhakar, B.S.; Meriggioli, M.N. Impaired regulatory function in circulating
CD4+CD25highCD127low/− T cells in patients with myasthenia gravis. Clin. Immunol. 2012, 145, 209–223. [CrossRef]
J. Clin. Med. 2021, 10, 2235 17 of 17

129. Mauri, C.; Menon, M. Human regulatory B cells in health and disease: Therapeutic potential. J. Clin. Investig. 2017, 127, 772–779.
[CrossRef]
130. Sheng, J.R.; Rezania, K.; Soliven, B. Impaired regulatory B cells in myasthenia gravis. J. Neuroimmunol. 2016, 297, 38–45. [CrossRef]
131. Sun, F.; Ladha, S.S.; Yang, L.; Liu, Q.; Bs, S.X.S.; Su, N.; Bomprezzi, R.; Shi, S.X.-Y. Interleukin-10 producing-B cells and their
association with responsiveness to rituximab in myasthenia gravis. Muscle Nerve 2013, 49, 487–494. [CrossRef] [PubMed]
132. Yi, J.S.; Russo, M.A.; Massey, J.M.; Juel, V.; Hobson-Webb, L.D.; Gable, K.; Raja, S.M.; Balderson, K.; Weinhold, K.J.; Guptill, J.T.
B10 Cell Frequencies and Suppressive Capacity in Myasthenia Gravis Are Associated with Disease Severity. Front. Neurol. 2017,
8, 34. [CrossRef] [PubMed]
133. Karim, R.; Zhang, H.-Y.; Yuan, J.; Sun, Q.; Wang, Y.-F. Regulatory B Cells in Seropositive Myasthenia Gravis versus Healthy
Controls. Front. Neurol. 2017, 8, 43. [CrossRef]
134. Thompson, J.S.; Bixler, S.A.; Qian, F.; Vora, K.; Scott, M.L.; Cachero, T.G.; Hession, C.; Schneider, P.; Sizing, I.D.; Mullen, C.; et al.
BAFF-R, a Newly Identified TNF Receptor That Specifically Interacts with BAFF. Science 2001, 293, 2108–2111. [CrossRef]
135. Ragheb, S.; Lisak, R.P. B-Cell-Activating Factor and Autoimmune Myasthenia Gravis. Autoimmune Dis. 2011, 2011, 1–10.
[CrossRef] [PubMed]
136. Kang, S.-Y.; Kang, C.-H.; Lee, K.-H. B-cell-activating factor is elevated in serum of patients with myasthenia gravis. Muscle Nerve
2016, 54, 1030–1033. [CrossRef] [PubMed]
137. Li, X.; Xiao, B.-G.; Xi, J.-Y.; Lu, C.-Z.; Lu, J.-H. Decrease of CD4+CD25highFoxp3+ regulatory T cells and elevation of CD19+BAFF-
R+ B cells and soluble ICAM-1 in myasthenia gravis. Clin. Immunol. 2008, 126, 180–188. [CrossRef]

You might also like